Skip to content
BY 4.0 license Open Access Published by De Gruyter May 3, 2023

Dynamic selection of visible wavelengths using resonant TiO2 nanostructures

  • Han-Don Um ORCID logo , Deokjae Choi , Amit Solanki , Emerald Huang , Kwanyong Seo and Fawwaz Habbal EMAIL logo
From the journal Nanophotonics

Abstract

All-dielectric nanoarrays have strong electromagnetic resonances with various interesting applications and are tuned by adjusting their geometrical parameters. However, their optical properties are permanently encoded during fabrication. This study presents robust dynamically tunable all-dielectric nanoresonators for controllable, reversible, and reproducible color filtering. Our design uses an array of TiO2 nanodiscs embedded in a transparent, stretchable polydimethylsiloxane (PDMS) membrane and exhibits a narrow spectral response due to Mie magnetic and electric dipole resonances hybridized with the TiO2 nanodiscs lattice modes. By mechanically stretching the PDMS membrane, the pitch of the TiO2 nanodiscs was increased and the spectral location of the resonances was altered. Additionally, an optically asymmetric structure was fabricated by partially embedding TiO2 nanodiscs in PDMS. Thus, the magnitude of the Rayleigh anomaly diffraction, which could interrupt the dipole resonances, was reduced. Our design has sharp, frequency-tunable resonances in the visible spectrum, and we demonstrated dynamic tunability by stretching the metasurfaces.

1 Introduction

Metasurfaces are optical materials with several innovative applications [18]. These materials have interesting light scattering properties, which enable the fabrication of small, lightweight optical devices, such as flat lenses [9], beam converters [10], deflectors [11], and holograms [12, 13]. Most dielectric metamaterials are arrays of subwavelength nanoresonators that strongly interact with the incident light [1419], producing resonances that are utilized in optical devices. These strong optical resonances are induced by Mie resonances, and enhanced by the collective periodicity of the nanoresonators lattice [20]. The observed resonances are an outcome of the coupling between two fundamentally different resonances: the Mie resonances and the lattice resonance. The Mie scattering is related to the polarizability of the material and its composition, as well as the shape of the nanoresonators. The lattice resonance is related to the periodicity of the nanoresonator array. The hybridization of these two resonances represents coupling between the relatively broad Mie resonance of individual scatterers and discrete states associated with the light scattered by the periodic array [2123]. This coupling is a function of several parameters such as the periodicity of the lattice structure and the optical properties of the surrounding medium.

The properties of the resonating elements and the lattice properties provide excellent opportunities for designing optical devices. Previous reports [2430] have demonstrated that the optical responses are tunable by modulating the geometric parameters of the array. However, these electromagnetic resonances are then permanently encoded by the array’s geometry, which is formed during fabrication. Recently, attempts have been made to tune the metasurface resonances using various materials, such as liquid crystals [3133], oxidized/reduced oxides [34], phase change materials [16], and solvents [35]. The dynamic modification of the metasurface performance using controllable, reversible, and reproducible methods is useful for various applications. Transparent and stretchable elastomers, such as polydimethylsiloxane (PDMS), have been successfully used as a substrate for tuning optical metasurfaces. By stretching the elastomer, the pitch (spacing) between the nanoresonators was changed to achieve different optical resonances [20].

This study presents a robust design for fabricating mechanically tunable all-dielectric nanoresonators for spectral filtering and other applications and shows conditions for observing both the electrical and magnetic resonances. A TiO2-based metasurface is fabricated using a procedure that is compatible with the industrially applicable complementary metal oxide semiconductor (CMOS) process [17].

Nanostructured TiO2 has a high index of refraction, n, larger than 2.5, and is suitable for imaging applications because it is transparent with insignificant optical losses in the visible spectrum [3]. A composite with a periodic square array of TiO2 nanodiscs embedded in a stretchable PDMS membrane is fabricated. By stretching the PDMS, the pitch between the TiO2 resonators changes, while the geometry of the hard TiO2 nanodiscs is maintained. Several TiO2 nanodiscs arrays are fabricated with dimensions appropriate to monitor the contributions of the electric and magnetic field components of the resonances and they exhibit sharp transmittance lines in the visible region of the electromagnetic spectrum.

It is recognized that the geometric parameters (diameter and height) of the array elements influence the resonances, and the collective behavior of the resonant elements affects the scattering properties and the corresponding resonances [1, 30]. Additionally, these collective electromagnetic resonances are sensitive to changes in the distance between the resonant nanostructures [36, 37]. This is a direct consequence of changes in the lattice resonances due to the change in the coupling between the adjacent resonant elements. When increasing the pitch, a longer wavelength of the light is needed to excite the same lattice resonances. Thus, spectral location of the resonances redshifts with increasing pitch. This dependence of the resonance wavelength on the pitch allows a dynamic tuning of the spectra.

Notably, there is another type of resonance that arises from the periodicity of the array. Any periodic scattering structure results in Rayleigh diffraction that may interact with the Mie lattice resonances, and even suppressing the lattice resonances. To clearly observe the dipole resonances, it is important to reduce the intensity of the Rayleigh diffraction by changing the index of the surrounding materials.

The optical characteristics of the lattice modes change with the pitch and the optical properties of the surrounding medium [38]. This was achieved by introducing optical asymmetry in the propagation direction of the incident light. With these changes, this study shows experimentally sharp optical resonances that are regulated by changing the pitch of the TiO2 nanoresonators and dynamically tuned in the visible spectrum. This type of spectral tuning is suitable for fabricating color filters that are dynamically selectable, which is useful for different imaging applications including multispectral spectroscopy.

2 Methods

2.1 Fabrication

Different PDMS-embedded TiO2 nanodisc arrays were prepared on quartz substrates. The quartz substrate was cleaned by sonication in acetone and isopropyl alcohol (IPA). The sacrificial germanium oxide layer was formed onto the quartz substrate by depositing a 100 nm-thick Ge layer using an e-beam evaporation system followed by annealing at 600 °C for 10 min under an O2 gas flow of 1000 sccm. This is a reliable method for fabricating a sacrificial germanium oxide layer. The TiO2 nanodisc arrays were formed using e-beam lithography followed by ALD deposition. To form the 275 nm thick electron-beam resist layer, a pristine electron-beam resist solution (ZEP520A, Zeon Corp.) was diluted with anisole solvent (99.7 %, Sigma-Aldrich) at a 1:2 volume ratio. The diluted e-beam resist solution was spin-coated onto the germanium oxide layer at a speed of 3000 rpm for 45 s. Subsequently, the substrate was annealed at 180 °C for 3 min. The desired hole pattern was written using e-beam lithography (ELS-125, Elionix) at a dose of 400 μC/cm2 and beam current of 1 nA, which produced the hole patterns over a 250 × 250 μm area. The hole diameter was 120 nm, and the pitch between the holes was 280 nm. The exposed e-beam resist was developed in anhydrous o-xylene (97 %, Sigma-Aldrich) for 45 s and then rinsed with IPA for 30 s. Subsequently, a 100 nm-thick TiO2 layer was deposited at 90 °C to fill the e-beam resist nanoholes. To remove the TiO2 layer that formed on top of the e-beam resist, the TiO2 surface was maintained at 20 °C and etched by the reactive-ion etching (RIE) process using BCl3 gas (40 sccm) and Ar gas (60 sccm), and source and stage powers of 500 W and 100 W, respectively. To prevent e-beam resist exposure, an O2 plasma treatment was performed under an O2 gas flow of 40 sccm at the room temperature, with a source power of 150 W and chamber pressure of 250 mTorr. To embed the TiO2 nanodisc array into PDMS, a mixed solution of PDMS base resin and curing agent (Sylgard 184, Dow Corning Co.) was spin-coated onto the TiO2 nanodisc array at a speed of 500 rpm for 3 min followed by annealing at 60 °C for 5 h. Subsequently, the cured PDMS with the TiO2 nanodisc array was immersed in water for 10 h to dissolve the germanium oxide layer, and finally the PDMS-embedded TiO2 nanodisc array was detached from the quartz substrate.

2.2 Optical characterization

Computer simulations of light propagating into the PDMS-embedded TiO2 nanodisc arrays were performed using a commercial software package (FDTD Solution, Lumerical Inc. and FDTD Band Solver, Optiwave systems Inc.) to obtain the transmission and reflection spectra. The refractive index of TiO2 used in the simulation was measured by ellipsometry (WVASE32, J.A. Woollam). For the optical simulation, a plane wave was injected onto the structure at a normal incidence with the electric field polarized along the x-axis. The x and y boundaries of the simulation domain were set with periodic boundary conditions. A perfectly matched layer at the z boundary of the domain was used, and simulations were performed with various unit cell sizes from 280 × 280 nm to 400 × 400 nm. Using a hyperspectral imaging microscope system (Horiba, XploRA), the reflection and transmission spectra of the samples were measured. To mechanically stretch the PDMS-embedded TiO2 nanodisc arrays, a homemade radial stretching system was designed. This radial stretching system was equipped with 6 arms that would serve to hold and stretch the sample. Each arm featured a peg that moves along a curved track on a rotating frame, allowing the arms to be pulled apart in a synchronized and controlled manner. The arms are guided within a fixed frame that is situated between two rotating frames. As a result, the sample dimensions can be increased from its original size by about 43 % (from 6 mm up to 8.5 mm).

3 Results and discussion

3.1 Key geometric parameter of TiO2 metasurface: pitch

To simulate the optical properties of the TiO2 metasurface, commercial software (Ansys Lumerical) was used to perform complete three-dimensional (3D) finite-difference time-domain (FDTD) analysis. First, an infinite array of TiO2 nanodiscs completely embedded in PDMS was considered, as shown in Figure 1(a), and the effects of changing the geometric parameters of the array were investigated. A unit cell of a single TiO2 nanodisc in the center with periodic boundary conditions in the x and y directions and a perfectly matched layer boundary in the z direction was used as the starting structure. A plane wave source with an electric field polarized along the x-axis and wave vector in the z direction was imposed onto the TiO2 nanodisc side and then propagated through PDMS. The reflectance and transmittance spectra were obtained by field monitors above and below the structure. The simulated optical response of the array was mapped by changing three geometrical parameters: nanodisc height (H), nanodisc diameter (D), and array pitch (P). Additionally, the refraction indices of the substrates and nanodiscs were used as simulation parameters. The refractive index of TiO2 used in the simulation was experimentally measured using a TiO2 thin film deposited by atomic layer deposition (ALD) (Supplementary Figure S1), which was also used to form the metamaterial TiO2 nanodiscs.

Figure 1: 
Simulation results of TiO2 nanodiscs completely embedded in uniform media with different indices. (a) Schematic of completely embedded TiO2 nanodiscs for the optical simulation. (b) Simulated transmittance spectra of TiO2 nanodiscs with different pitches and heights, at a fixed diameter of 120 nm. (c) Simulated transmittance spectra of TiO2 nanodiscs with different pitches and diameters, at a fixed height of 275 nm. (d) Transmittances versus pitch of the completely embedded TiO2 nanodiscs with a diameter of 120 nm and thickness of 275 nm.
Figure 1:

Simulation results of TiO2 nanodiscs completely embedded in uniform media with different indices. (a) Schematic of completely embedded TiO2 nanodiscs for the optical simulation. (b) Simulated transmittance spectra of TiO2 nanodiscs with different pitches and heights, at a fixed diameter of 120 nm. (c) Simulated transmittance spectra of TiO2 nanodiscs with different pitches and diameters, at a fixed height of 275 nm. (d) Transmittances versus pitch of the completely embedded TiO2 nanodiscs with a diameter of 120 nm and thickness of 275 nm.

The transmission resonances were simulated for different heights at a fixed diameter. This was repeated for different pitches, as shown in Figure 1(b). The heights were 225, 275, and 325 nm, and pitches were 280, 340, and 400 nm, while the diameter was fixed at 120 nm. The transmittance spectra exhibited sharp decrease at particular spectral locations (dips), which did not significantly shift with changes in H in the range of 275 ± 50 nm. In contrast, the dips significantly shifted with increasing array pitch from 280 to 400 nm. Subsequently, the diameter was varied (D = 110, 120, and 130 nm) at a fixed height of 275 nm, and the simulations were repeated at different pitches. The transmittance spectra are shown in Figure 1(c). From Figure 1(b) and (c), the location of the spectral dips depended strongly on the array pitch and weakly on the diameter and height of the nanodiscs. To evaluate the spectral dependence on pitch, a simulation was performed using a fine pitch sweep with P = 280–500 nm for nanodiscs with a diameter of 120 nm and height of 275 nm, as shown in Figure 1(d). The spectral location of the transmission bands exhibited redshifts with increasing pitch, and crossover at λ ∼ 295 nm. These results are consistent with data in the literature [29]. The locations of the scattering modes are controlled by the size and refractive index of the nanostructures, i.e., the Mie scattering cross sections, and it is essential to have high index material to exhibit resonances in the visible wavelengths. On the level of the resonating elements, the electric dipoles scatter light symmetrically in the plane transverse to the dipole axis, and the resonant magnetic dipole results from a coupling of incoming light to the circular displacement currents of the electric field, due to the field penetration and phase retardation inside the nanodiscs. It is well known that the geometric parameters (diameter and height) of the elements of the array influence the resonances. The collective optical resonances are sensitive to the change in the distance between the resonant nanostructures [36, 37]. This is a direct consequence to changes in the coupling between adjacent elements. This electromagnetic coupling enhances with decreasing the pitch [37], and in a large array of resonators, the nearest neighbors provide the strongest contribution to the overall resonances. The electromagnetic coupling is sensitive to changes in pitch between the scattering centers, and reducing this coupling leads to changes in resonance intensities [24, 39]. When the spacing between the scattering elements increased, the coupling is significantly reduced and then diminished, and the resonance spectra disappeared (Figure 1(d)). One of the resonances diminished at pitch values higher than ∼385 nm, and further increasing the pitch above 545 nm weakened the other resonance (Supplementary Figure S2). Thus, stretching PDMS can tune the transmission spectra because coupling is the main driver for spectral tuning.

3.2 Reducing the Rayleigh Anomaly diffraction

To elucidate the origin of the electromagnetic resonances A and B in Figure 1(d), FDTD simulations were used to obtain the electric and magnetic field distributions within the nanodiscs. For example, the field distributions for nanodiscs with P = 330 nm are shown in Figure 2(a)–(c). For resonance A, the electric field aligned along the direction of incident polarization (Figure 2(a) and (b)), whereas the magnetic field vector curls in the cross section (YZ plane) shown in Figure 2(c). For resonance B, the electric field vector curls in the cross section (XZ plane) shown in Figure 2(e), and the magnetic field vector was perpendicular to the direction of incident polarization (Figure 2(d) and (f)). Therefore, A is the electric dipole resonance and B is the magnetic dipole resonance. Notably, when the resonances shifted with decreasing spacing, crossover of the electric dipole resonance was observed at shorter wavelengths than that of the magnetic dipole resonance (Supplementary Figure S3). In Figure 1(d), the Rayleigh anomaly (RA) diffraction was observed close to the Mie-lattice resonances [22, 40], [41], [42]. This diffraction originates from the periodic structure that serves as a thin transmission grating and diffracts the impinging light [43]. To utilize the Mie-lattice resonances for practical applications, it is essential to reduce the RA intensity. Therefore, we added another parameter in the design of the array by creating an optically asymmetric structure which affects the RA.

Figure 2: 
Electric and magnetic field intensity plotted along the vertical TiO2 nanodisc: (a)–(c) electric dipole resonance at λ = 502 nm and (d)–(f) magnetic dipole resonance at λ = 495 nm for the completely embedded TiO2 nanodisc array with a pitch of 330 nm.
Figure 2:

Electric and magnetic field intensity plotted along the vertical TiO2 nanodisc: (a)–(c) electric dipole resonance at λ = 502 nm and (d)–(f) magnetic dipole resonance at λ = 495 nm for the completely embedded TiO2 nanodisc array with a pitch of 330 nm.

The general condition for the RA diffraction was described by Strutt [40].

(1) n sub S i n β m = n sup S i n α inc ± m λ / P

where n sub is the refractive index of the substrate, β m is the diffraction angle in the substrate, n sup is the refractive index of the superstrate, a inc is the light angle in the superstrate, integer m is the order of the diffracted wave, λ is the wavelength of the light, and P is the pitch of the diffracting elements (the TiO2 nanodiscs). Because the incident light is normal to the nanodisc array (along the z axis), a inc is 0° and n sub and n sup are identical. When the angle β m is 90°, the RA occurs due to the propagation of the diffracted wave parallel to the grating. Thus, for the geometry of our system, the first order diffraction grating Equation (1) becomes:

(2) λ = n sub × P ,

Equation (2) indicates that changing the optical index of the embedding medium shifts the spectral location of the RA resonance. For example, reducing the refractive index of the medium caused a blueshift in the RA resonance (Supplementary Figure S4), but the change in the index also shifts the dipole resonances. Subsequently, a metasurface with shifted RA diffraction was designed, which consists of an optically asymmetric structure in which the top and bottom surfaces of the TiO2 nanodiscs are in contact with two media with different refractive indices (for example, PDMS and air). This structure was realized using partially embedded nanodiscs in PDMS. Figure 3(a) and (b) shows transmission and reflection when light was injected from the PDMS side, exhibiting a strong RA reflection. Figure 3(c) and (d) were obtained when light was injected from the air side, exhibiting a weak RA reflection. These results are attributed to the change in index at the interfaces. Because the effective index for the TiO2-PDMS composite is approximately 1.6 and that of air is 1, when light is injected from the PDMS side and propagates through air, the index is significantly reduced (Δn = 0.6), which produces a strong reflection at the interface between the TiO2-PDMS composite and air. Conversely, when light is injected from the air side, a smaller decrease in the index was observed Δn = 1.6–1.46 (the index of SiO2) = 0.14, causing a weaker RA reflection.

Figure 3: 
Asymmetric media structure of partially embedded TiO2 nanodiscs in PDMS with a diameter of 120 nm and thickness of 275 nm for Rayleigh anomaly resonance. (a) Transmittance and (b) reflectance spectra versus pitch of the TiO2 nanodiscs with light injected from the PDMS side and transmitted through air. (c) Transmittance and (d) reflectance spectra versus pitch of the TiO2 nanodiscs with light injected from air and transmitted through PDMS. Insets of panel (c) and (d) show the schematics of simulated TiO2 nanodisc structures.
Figure 3:

Asymmetric media structure of partially embedded TiO2 nanodiscs in PDMS with a diameter of 120 nm and thickness of 275 nm for Rayleigh anomaly resonance. (a) Transmittance and (b) reflectance spectra versus pitch of the TiO2 nanodiscs with light injected from the PDMS side and transmitted through air. (c) Transmittance and (d) reflectance spectra versus pitch of the TiO2 nanodiscs with light injected from air and transmitted through PDMS. Insets of panel (c) and (d) show the schematics of simulated TiO2 nanodisc structures.

3.3 Fabrication of dynamically tunable TiO2 metasurface

The dynamic tuning of the TiO2 nanodisc color filtering system was realized experimentally by mechanically stretching PDMS with the partially embedded TiO2 nanodiscs. Stamping methods had been used to transfer nanostructures into flexible substrates [44, 45]. These transfers may form defects in the nanopattern, which reduces device fidelity. As shown in Figure 4, a different fabrication method for embedding TiO2 nanodiscs in PDMS was designed to produce defect-free structures. A sacrificial layer of GeO x film, which can be dissolved in water, was deposited onto a glass substrate. This layer was coated with an electron-beam resist film and patterned by electron-beam lithography to produce a nanohole array. Subsequently, the nanoholes were filled with TiO2 using ALD, which has excellent step coverage. Near-perfect coverage was confirmed by evaluating a filled nano-trench (Figure S5). As expected, filling the nanoholes resulted in a thin TiO2 film on the electron-beam resist, which was completely removed by reactive ion etching. Thus, the height of TiO2 nanodiscs was determined by the electron-beam resist thickness. Therefore, the electron-beam resist was selectively etched by O2 plasma treatment to form a TiO2 nanodisc array on the GeOx layer. The use of a dry lift-off step provided a defect-free TiO2 metasurface. Subsequently, a PDMS solution was poured onto the TiO2 nanodiscs and cured. Finally, the device was immersed in water to remove the GeO x layer. The combination of the lift-off step and sacrificial layer produces a defect-free stretchable metasurface. The scanning electron microscopy (SEM) image of as-fabricated TiO2 nanodiscs is shown in Figure S6.

Figure 4: 
Fabrication process of PDMS-embedded TiO2 nanodisc array-based color filters.
Figure 4:

Fabrication process of PDMS-embedded TiO2 nanodisc array-based color filters.

3.4 Experimental demonstration of magnetic dipole resonances

The transmittance spectra of the metasurface were measured using a commercial hyperspectral imaging system. By mechanically stretching the PDMS, the pitch between the TiO2 nanodiscs was dynamically changed from 280 to 380 nm. Figure 5(a) shows spectra redshift with increasing pitch similar to the simulation results in Figure 1(d). For comparison, Figure 5(b) shows the simulated spectra of the metasurface dimensions, which corresponds with the experimental results, exhibiting a similar redshift trend. However, the simulated bands were much narrower with transmission minima of <10 % than the experiment ones, which exhibited broader transmission bands with minima of >30 %. Additionally, a single dip was observed in the experimental spectra, whereas the simulated spectra showed double dips representing the electric and magnetic dipole resonances. These differences are commonly reported for dielectric metasurface arrays [29, 30, 35] but the fact that often only one broad dip is observed remains unresolved. It is crucial to understand the origin of these differences to achieve the theoretical near-zero transmissions and be able to obtain high-quality resonances.

Figure 5: 
Dynamic color tuning by stretching PDMS-embedded TiO2 nanodiscs: (a) experimental and (b) simulated transmittance spectra of PDMS-embedded TiO2 nanodiscs with different pitches from 280 to 380 nm. The Insets in panel (a) show the optical microscopy images of PDMS-embedded TiO2 nanodiscs with different pitches from 280 to 380 nm. Transmittance spectra (c) and (d) for PDMS-embedded TiO2 nanodiscs and (e) and (f) for PMMA-coated TiO2 nanodiscs on glass substrate at a pitch of 280 and 340 nm – non-averaged simulated spectrum (blue line), simulated spectrum from averaged spectra with various angles of incident light (red line), and experimental spectrum (green line). The insets in panel (c)–(f) show the schematics of the TiO2 nanodiscs with different media.
Figure 5:

Dynamic color tuning by stretching PDMS-embedded TiO2 nanodiscs: (a) experimental and (b) simulated transmittance spectra of PDMS-embedded TiO2 nanodiscs with different pitches from 280 to 380 nm. The Insets in panel (a) show the optical microscopy images of PDMS-embedded TiO2 nanodiscs with different pitches from 280 to 380 nm. Transmittance spectra (c) and (d) for PDMS-embedded TiO2 nanodiscs and (e) and (f) for PMMA-coated TiO2 nanodiscs on glass substrate at a pitch of 280 and 340 nm – non-averaged simulated spectrum (blue line), simulated spectrum from averaged spectra with various angles of incident light (red line), and experimental spectrum (green line). The insets in panel (c)–(f) show the schematics of the TiO2 nanodiscs with different media.

We suggest two points to investigate: (1) the origin of the differences between the computer simulations and experimental setups, and (2) the effect of the presence of slight variations in the array pitch and also the possibility of surface non-uniformity caused by the PDMS transfer or handling. For the former, an ideal case with a plane wave injection at a normal incidence was used in the simulations, whereas a light injection angle set by the numerical aperture (NA) of the objective lens was used experimentally. The experimental objective lens had NA = 0.13, which corresponds to a half angle of incidence equivalent to ±7.4° relative to the normal incidence. It is recognized that the transmission spectra of periodic dielectric nanostructures are dependent on the incidence angle [17, 24]. When the transmission spectra are averaged over the incidence angle, the resonance band broadening is reasonable. This was investigated by performing FDTD simulations in the incident angle range of 82.6°–97.4° with 0.3° steps. The simulated spectra with different incident angles are shown in Figure S7. A shift in the resonance position was observed with changes in polar angle (θ) and azimuthal angle (φ). The simulated data were averaged and plotted (in red) in Figure 5(c) with the transmission spectrum of the ideal case (plotted in blue) at a normal incidence for PDMS-embedded TiO2 nanodiscs with P = 280 nm. The averaged spectra exhibited a much less transmission minimum (38 %) than that at a normal incidence which were about 8 %. Notably, the averaged spectra exhibited overlapping spectra dips, which correspond with the experimental spectrum (Figure 5(c), plotted in green). A similar spectrum was obtained at P = 340 nm as shown in Figure 5(d). The transmission minimum is reduced for the angle averaged simulated plots compared with the normal incidence simulated plots, as shown in Figure 5(c) and (d). The weaker spectra dips in Figure 5(a) are due to the NA (0.13) of the objective lens in experiment compared with the normal incidence simulation in Figure 5(b).

The second hypothesis, which attributes the differences between experimental and simulated results to slight variations in the spatial dimensions during device fabrication was also investigated. Spatial variations and surface non-uniformity can occur during the (a) transfer of the TiO2 metasurface to PDMS or (b) mounting of the thin flexible metasurface onto the mechanical stretching mount. Furthermore, these non-uniformities can be amplified by the stretching process. A small local variation in the position of the array elements may cause incoherency in strongly coupled modes, resulting in broader collective resonance. To investigate this hypothesis, TiO2 nanodiscs were formed on a rigid glass substrate to eliminate bending, twisting, or variations in the pitch between the nanodiscs during the transfer process. A structure similar to PDMS-embedded TiO2 metasurface was implemented by fabricating the TiO2 metasurface on a glass substrate and embedding the metasurface in a polymethyl methacrylate (PMMA) acrylic glass with a thickness equal to the nanodisc height. The refractive index of PMMA is approximately 1.48, which is close to that of PDMS (n = 1.46), and thus the two systems are similar. The low viscosity of the PMMA solution enabled easy filling between the TiO2 nanodiscs by spin coating. The dipole resonances of the PMMA-containing structure were simulated at a normal incidence and incident light with different angles. Figure 5(e) shows the simulated transmittance spectrum at a normal incidence (blue) and averaged transmittance spectra with different angles (red) for the PMMA-coated TiO2 metasurface on the glass substrate (P = 280 nm). The averaged spectra of the glass/PMMA metasurface exhibited a similar trend with overlapping bands. It was experimentally confirmed that the spectra (green in Figure 5(e)) for the glass/PMMA metasurface at P = 280 nm resembles the simulation spectra obtained with illumination angle averaged (red in Figure 5(e)) for this geometry. Notably, for the glass/PMMA metasurface (green in Figure 5(e)), the double transmission dips were experimentally observed and resemble those in the simulation. In addition, the measured dip was sharper with a full width at half maximum (FWHM) of 7.2 nm, compared to the case of metasurface embedded in PDMS (green in Figure 5(c)) with FWHM = 9.0 nm. Subsequently, as shown in Figure 5(f), the pitch was increased to 340 nm, and the experimental data (in green) for the electric and magnetic dipole resonances show two distinct dips which are similar to the case for the normal incidence light in the averaged simulated spectra. Based on these results, we concluded that the illumination angle variations strongly influence the intensities of the transmission dips. Moreover, the spatial non-uniformity induced in the metasurface during the mechanical stretching causes band broadening due to pitch variations. The measured FWHM of the spectra of the glass/PMMA metasurface is 3.4 nm (green in Figure 5(f)) was larger than that of the averaged simulation (red in Figure 5(f)), which has FWHM = 2.1 nm.

According to Figure 5(c)–(f), we concluded that pitch variation during mechanically stretching of the PDMS is likely more dominant for broadening the experimental spectra. However, these variations do not negatively affect the functionality of the metasurface as a narrow-band tunable filter for the dynamic selection of visible wavelengths. To accurately control the pitch of the PDMS-embedded TiO2 nanodiscs, a stretching system is being developed using dielectric elastomer actuators [46] instead of a mechanical force. Using the dielectric elastomer actuators, it is possible to stretch along the radial direction and control the stretching electrically. Thus, a more uniform and accurate control of nanodisc array pitch is expected, which may produce narrow-band TiO2 nanodisc color filters.

4 Conclusions

This study presents a reliable method for fabricating dynamically tuned TiO2-based dielectric metasurfaces using a mechanically stretchable elastomer. Dynamically tunable color filters were demonstrated by mechanically stretching the PDMS-embedded dielectric metasurface up to 43 % and regulating the spectral location of the Mie lattice resonances by changing the pitch of the array. Moreover, sharp, and variable wavelength spectra transmission dips in the range of 435–580 nm were achieved. Additionally, the Rayleigh anomaly diffraction was shifted by embedding TiO2 nanostructures in an optically asymmetric structure enabling a more flexible engineering of transmission filters.

From simulations and experiments we showed that the non-uniformity of the stretching can increase the broadness of the lattice resonances, while focusing the incident light through large NA optics leads to averaging the transmittance spectra and reduces the quality factor.

Due to the high transparency of TiO2 and the PDMS in the visible region, the PDMS-embedded TiO2 dielectric metasurfaces provide lossless resonances. With a process compatible with the industrial CMOS fabrication process, we created defect-free photonic dielectric metasurfaces that can be used as narrow band filters in the visible light. The tuning region of the developed device coincides with the spectral region of visible light cameras. Thus, by coupling these filters with appropriate photodetectors, they can be used in hyperspectral imaging and other sensing applications.


Corresponding author: Fawwaz Habbal, John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA, E-mail: habbalf@seas.harvard.edu

Han-Don Um and Deokjae Choi are equal contributors.


Funding source: National Research Foundation of Korea

Award Identifier / Grant number: NRF-2019R1A2C2086602, NRF-2019M1A2A2065614, NRF-2021R1F1A1064243, NRF-2022R1C1C1010025, NRF-2022R1C1C2004320

Funding source: Harvard University

Funding source: National Science Foundation

Award Identifier / Grant number: NSF ECCS award no. 1541959

Funding source: Health@InnoHK program of the Innovation and Technology Commission of Hong Kong.

  1. Author contribution: H.D.U., D.C., and F.H. conceived and designed the research. H.D.U., D.C., E.H., and A.S. performed the experiments and the optical simulations. H.D.U., A. S., K.S., and F.H. wrote the manuscript.

  2. Research funding: This work was performed in part at the Harvard University Center for Nanoscale Systems (CNS), a member of the National Nanotechnology Coordinated Infrastructure Network (NNCI), which is supported by the National Science Foundation under NSF ECCS award no. 1541959. It was also partially supported by the National Research Foundation of Korea (NRF) grant funded by the Korean Government (MSIP) (Grants NRF-2019R1A2C2086602, NRF-2019M1A2A2065614, NRF-2021R1F1A1064243, NRF-2022R1C1C1010025, and NRF-2022R1C1C2004320), and by the Health@InnoHK program of the Innovation and Technology Commission of Hong Kong.

  3. Conflict of interest statement: The authors declare no competing financial interest.

References

[1] W. Yang, S. Xiao, Q. Song, et al.., “All-dielectric metasurface for high-performance structural color,” Nat. Commun., vol. 11, no. 1, p. 1864, 2020. https://doi.org/10.1038/s41467-020-15773-0.Search in Google Scholar PubMed PubMed Central

[2] G. W. Castellanos, S. Murai, T. V. Raziman, et al.., “Exciton-polaritons with magnetic and electric character in all-dielectric metasurfaces,” ACS Photonics, vol. 7, p. 50, 2020. https://doi.org/10.1021/acsphotonics.0c00063.Search in Google Scholar

[3] Y. Wu, W. Yang, Y. Fan, et al.., “TiO2 metasurfaces: from visible planar photonics to photochemistry,” Sci. Adv., vol. 5, no. 11, p. eaax0939, 2019. https://doi.org/10.1126/sciadv.aax0939.Search in Google Scholar PubMed PubMed Central

[4] Y. Fan, Y. Wang, N. Zhang, et al.., “Resonance-enhanced three-photon luminesce via lead halide perovskite metasurfaces for optical encoding,” Nat. Commun., vol. 10, no. 1, pp. 1–8, 2019. https://doi.org/10.1038/s41467-019-10090-7.Search in Google Scholar PubMed PubMed Central

[5] L. Jin, Y. W. Huang, Z. Jin, et al.., “Dielectric multi-momentum meta-transformer in the visible,” Nat. Commun., vol. 10, no. 1, pp. 1–6, 2019. https://doi.org/10.1038/s41467-019-12637-0.Search in Google Scholar PubMed PubMed Central

[6] A. Tittl, A. Leitis, M. Liu, et al.., “Imaging-based molecular barcoding with pixelated dielectric metasurfaces,” Science, vol. 360, no. 6393, pp. 1105–1109, 2018. https://doi.org/10.1126/science.aas9768.Search in Google Scholar PubMed

[7] Y. Gao, Y. Fan, Y. Wang, et al.., “Nonlinear holographic all-dielectric metasurfaces,” Nano Lett., vol. 18, no. 12, pp. 8054–8061, 2018. https://doi.org/10.1021/acs.nanolett.8b04311.Search in Google Scholar PubMed

[8] A. D. Boardman, V. V. Grimalsky, Y. S. Kivshar, et al.., “Active and tunable metamaterials,” Laser Photon. Rev., vol. 5, no. 2, pp. 287–307, 2011. https://doi.org/10.1002/LPOR.201000012.Search in Google Scholar

[9] H.-S. Ee and R. Agarwal, “Tunable metasurface and flat optical zoom lens on a stretchable substrate,” Nano Lett., vol. 16, p. 44, 2016. https://doi.org/10.1021/acs.nanolett.6b00618.Search in Google Scholar PubMed

[10] F. Alves, L. Pimental, D. Grbovic, et al.., “MEMS terahertz-to-infrared band converter using frequency selective planar metamaterial,” Sci. Rep., vol. 8, no. 1, p. 12466, 2018. https://doi.org/10.1038/s41598-018-30858-z.Search in Google Scholar PubMed PubMed Central

[11] Z. Liu, Z. Li, Z. Liu, et al.., “High-performance broadband circularly polarized beam deflector by mirror effect of multinanorod metasurfaces,” Adv. Funct. Mater., vol. 25, no. 34, pp. 5428–5434, 2015. https://doi.org/10.1002/adfm.201502046.Search in Google Scholar

[12] G. Zheng, H. Mühlenbernd, M. Kenney, et al.., “Metasurface holograms reaching 80% efficiency,” Nat. Nanotechnol., vol. 10, no. 4, pp. 308–312, 2015. https://doi.org/10.1038/nnano.2015.2.Search in Google Scholar PubMed

[13] L. Wang, W. Zhang, H. Yin, et al.., “Ultrasmall optical vortex knots generated by spin-selective metasurface holograms,” Adv. Opt. Mater., vol. 7, no. 10, p. 1900263, 2019. https://doi.org/10.1002/adom.201900263.Search in Google Scholar

[14] M. Decker, I. Staude, M. Falkner, et al.., “High-Efficiency dielectric Huygens’ surfaces,” Adv. Opt. Mater., vol. 3, no. 6, pp. 813–820, 2015. https://doi.org/10.1002/adom.201400584.Search in Google Scholar

[15] M. Decker, T. Pertsch, and I. Staude, “Strong coupling in hybrid metal-dielectric nanoresonators,” Phil. Trans. R. Soc. A, vol. 375, no. 2090, p. 20160312, 2017. https://doi.org/10.1098/rsta.2016.0312.Search in Google Scholar PubMed PubMed Central

[16] A. Leitis, A. Heßler, S. Wahl, et al., “All-dielectric programmable Huygens’ metasurfaces,” Adv. Funct. Mater., vol. 30, no. 19, p. 1910259, 2020. https://doi.org/10.1002/adfm.201910259.Search in Google Scholar

[17] A. I. Kuznetsov, A. E. Miroshnichenko, M. L. Brongersma, et al.., “Optically resonant dielectric nanostructures,” Science, vol. 354, no. 6314, p. aag2472, 2016. https://doi.org/10.1126/science.aag2472.Search in Google Scholar PubMed

[18] L. Lewin, “The electrical constants of a material loaded with spherical particles,” J. Inst. Electr. Eng. - Part III Radio Commun. Eng., vol. 94, no. 27, pp. 65–68, 1947. https://doi.org/10.1049/JI-3-2.1947.0013.Search in Google Scholar

[19] C. F. Bohren and D. R. Huffman, “Absorption and scattering of light by small particles,” in Absorpt. Scatt. Light by Small Part, 1998.10.1002/9783527618156Search in Google Scholar

[20] A. B. Evlyukhin, S. M. Novikov, U. Zywietz, et al.., “Demonstration of magnetic dipole resonances of dielectric nanospheres in the visible region,” Nano Lett., vol. 12, no. 7, pp. 3749–3755, 2012. https://doi.org/10.1021/NL301594S/ASSET/IMAGES/LARGE/NL-2012-01594S_0008.JPEG.Search in Google Scholar

[21] J. A. Garcia, C. Hrelescu, X. Zhang, et al.., “Quasi-guided modes in titanium dioxide arrays fabricated via soft nanoimprint lithography,” ACS Appl. Mater. Interfaces, vol. 13, no. 40, pp. 47860–47870, 2021. https://doi.org/10.1021/ACSAMI.1C11456/ASSET/IMAGES/MEDIUM/AM1C11456_M006.GIF.Search in Google Scholar

[22] M. Kim, E. S. Hwang, O. Prudnikov, et al.., “Guided-mode waves structure of electric and magnetic dipole resonances in a metamaterial slab,” Opt. Express, vol. 29, no. 19, pp. 30610–30624, 2021. https://doi.org/10.1364/OE.437899.Search in Google Scholar PubMed

[23] Y. Li, X. Huang, S. Huang, et al.., “Tunable dual-band perfect metamaterial absorber based on monolayer graphene arrays as refractive index sensor,” Jpn. J. Appl. Phys., vol. 59, no. 9, p. 095002, 2020. https://doi.org/10.35848/1347-4065/ABA9A5.Search in Google Scholar

[24] V. E. Babicheva and A. B. Evlyukhin, “Resonant lattice kerker effect in metasurfaces with electric and magnetic optical responses,” Laser Photon. Rev., vol. 11, no. 6, p. 1700132, 2017. https://doi.org/10.1002/lpor.201700132.Search in Google Scholar

[25] Y. H. Ko and R. Magnusson, “Wideband dielectric metamaterial reflectors: Mie scattering or leaky Bloch mode resonance?” Optica, vol. 5, no. 3, p. 289, 2018. https://doi.org/10.1364/optica.5.000289.Search in Google Scholar

[26] B. Sain, C. Meier, and T. Zentgraf, “Nonlinear optics in all-dielectric nanoantennas and metasurfaces: a review,” Adv. Photonics, vol. 1, no. 02, p. 1, 2019. https://doi.org/10.1117/1.ap.1.2.024002.Search in Google Scholar

[27] G. W. Castellanos, P. Bai, and J. Gómez Rivas, “Lattice resonances in dielectric metasurfaces,” J. Appl. Phys., vol. 125, no. 21, p. 213105, 2019. https://doi.org/10.1063/1.5094122.Search in Google Scholar

[28] J. C. Ginn, I. Brener, D. W. Peters, et al., “Realizing optical magnetism from dielectric metamaterials,” Phys. Rev. Lett., vol. 108, no. 9, p. 097402, 2012. https://doi.org/10.1103/PhysRevLett.108.097402.Search in Google Scholar PubMed

[29] P. Gutruf, C. Zou, W. Withayachumnankul, et al.., “Mechanically tunable dielectric resonator metasurfaces at visible frequencies,” ACS Nano, vol. 10, no. 1, pp. 133–141, 2016. https://doi.org/10.1021/acsnano.5b05954.Search in Google Scholar PubMed

[30] C. Zhang, J. Jing, Y. Wu, et al.., “Stretchable all-dielectric metasurfaces with polarization-insensitive and full-spectrum response,” ACS Nano, vol. 14, no. 2, pp. 1418–1426, 2020. https://doi.org/10.1021/acsnano.9b08228.Search in Google Scholar PubMed

[31] A. Komar, Z. Fang, J. Bohn, et al.., “Electrically tunable all-dielectric optical metasurfaces based on liquid crystals,” Appl. Phys. Lett., vol. 110, no. 7, p. 071109, 2017. https://doi.org/10.1063/1.4976504.Search in Google Scholar

[32] J. Sautter, I. Staude, M. Decker, et al.., “Active tuning of all-dielectric metasurfaces,” ACS Nano, vol. 9, no. 4, pp. 4308–4315, 2015. https://doi.org/10.1021/ACSNANO.5B00723/ASSET/IMAGES/LARGE/NN-2015-00723C_0006.JPEG.Search in Google Scholar

[33] J. Bohn, T. Bucher, K. E. Chong, et al.., “Active tuning of spontaneous emission by mie-resonant dielectric metasurfaces,” Nano Lett., vol. 18, no. 6, pp. 3461–3465, 2018. https://doi.org/10.1021/acs.nanolett.8b00475.Search in Google Scholar PubMed

[34] X. Duan, S. Kamin, and N. Liu, “Dynamic plasmonic colour display,” Nat. Commun., vol. 8, no. 1, pp. 1–9, 2017. https://doi.org/10.1038/ncomms14606.Search in Google Scholar PubMed PubMed Central

[35] S. Sun, W. Yang, C. Zhang, et al., “Real-time tunable colors from microfluidic reconfigurable all-dielectric metasurfaces article,” ACS Nano, vol. 12, no. 3, pp. 2151–2159, 2018. https://doi.org/10.1021/acsnano.7b07121.Search in Google Scholar PubMed

[36] E. J. Heller, “Quantum proximity resonances,” Phys. Rev. Lett., vol. 77, no. 20, pp. 4122–4125, 1996. https://doi.org/10.1103/PhysRevLett.77.4122.Search in Google Scholar PubMed

[37] D. A. Powell, M. Lapine, M. V. Gorkunov, et al.., “Metamaterial tuning by manipulation of near-field interaction,” Phys. Rev. B - Condens. Matter Mater. Phys., vol. 82, no. 15, p. 155128, 2010. https://doi.org/10.1103/PhysRevB.82.155128.Search in Google Scholar

[38] J. Jang, T. Badloe, Y. Yang, et al.., “Spectral modulation through the hybridization of Mie-scatterers and quasi-guided mode resonances: realizing full and gradients of structural color,” ACS Nano, vol. 14, no. 11, pp. 15317–15326, 2020. https://doi.org/10.1021/ACSNANO.0C05656/ASSET/IMAGES/MEDIUM/NN0C05656_M010.GIF.Search in Google Scholar

[39] J. Li, N. Verellen, and P. Van Dorpe, “Engineering electric and magnetic dipole coupling in arrays of dielectric nanoparticles,” J. Appl. Phys., vol. 123, no. 8, p. 083101, 2018. https://doi.org/10.1063/1.5018312.Search in Google Scholar

[40] J. W. Strutt, “XII. On the manufacture and theory of diffraction gratings,” Philos. Mag., vol. 47, pp. 81–93, 1874, https://doi.org/10.1080/14786447408640996.Search in Google Scholar

[41] R. W. Wood, “On a remarkable case of uneven distribution of light in a diffraction grating spectrum,” Philos. Mag., vol. 4, pp. 396–402, 1902, https://doi.org/10.1088/1478-7814/18/1/325.Search in Google Scholar

[42] K. Baek, Y. Kim, S. Mohd-Noor, et al.., “Mie resonant structural colors,” ACS Appl. Mater. Interfaces, vol. 12, no. 5, pp. 5300–5318, 2020.10.1021/acsami.9b16683Search in Google Scholar PubMed

[43] S. Hamdad, A. T. Diallo, M. Chakaroun, et al.., “The role of Rayleigh anomalies in the coupling process of plasmonic gratings and the control of the emission properties of organic molecules,” Sci. Rep., vol. 12, no. 1, pp. 1–11, 2022. https://doi.org/10.1038/s41598-022-07216-1.Search in Google Scholar PubMed PubMed Central

[44] M. Zhu and C. Lee, “Facile metal transfer method for fabricating unconventional metamaterial devices,” Opt. Mater. Express, vol. 5, no. 4, p. 733, 2015. https://doi.org/10.1364/ome.5.000733.Search in Google Scholar

[45] J. W. Jeong, S. R. Yang, Y. H. Hur, et al.., “High-resolution nanotransfer printing applicable to diverse surfaces via interface-targeted adhesion switching,” Nat. Commun., vol. 5, no. 1, pp. 1–12, 2014. https://doi.org/10.1038/ncomms6387.Search in Google Scholar PubMed

[46] M. Bozlar, C. Punckt, S. Korkut, et al.., “Dielectric elastomer actuators with elastomeric electrodes,” Appl. Phys. Lett., vol. 101, no. 9, p. 091907, 2012. https://doi.org/10.1063/1.4748114.Search in Google Scholar


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/nanoph-2023-0057).


Received: 2023-01-29
Accepted: 2023-04-21
Published Online: 2023-05-03

© 2023 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2023-0057/html
Scroll to top button